• Aucun résultat trouvé

Electron doping and superconductivity in the two-dimensional Hubbard model

N/A
N/A
Protected

Academic year: 2022

Partager "Electron doping and superconductivity in the two-dimensional Hubbard model"

Copied!
4
0
0

Texte intégral

(1)

Electron doping and superconductivity in the two-dimensional Hubbard model

D. Eichenberger and D. Baeriswyl

Department of Physics, University of Fribourg, CH-1700 Fribourg, Switzerland

An elaborate variational wave function is used for studying superconductivity in the 共repulsive兲 two- dimensional Hubbard model, including both nearest- and next-nearest-neighbor hoppings. A marked asymme- try is found between the “localized” hole-doped region and the more itinerant electron-doped region. Super- conductivity withd-wave symmetry turns out to be restricted to densities where the Fermi surface crosses the magnetic zone boundary. A concomitant peak in the magnetic structure factor at 共␲,␲兲 clearly points to a magnetic mechanism.

One of the central issues in the field of high-temperature superconductors has been—and for some researchers still is—the question whether pairing in the cuprates arises from purely repulsive interactions, as proposed by Anderson two decades ago.1This question has been studied extensively in the framework of the two-dimensional共2D兲 共repulsive兲Hub- bard model 共and the related t-J model兲. Recent progress, both in dynamical mean-field theory2,3 and in variational calculations,4has strengthened the case for the existence of a superconducting phase in the Hubbard model, with a 共d-wave兲gap parameter reasonably close to the experimental values for intermediate interaction strengths 共Uof the order of the bandwidth兲. This conclusion has been challenged on the basis of Monte Carlo simulations,5which we believe to be not conclusive, as discussed below.

Most previous studies of the Hubbard model have been restricted to nearest-neighbor hopping, where electron dop- ing does not differ from hole doping. Here we show that the addition of second-neighbor hopping 共which breaks the electron-hole symmetry兲changes this behavior substantially.

While the hole-doped side is “localized” and shows kinetic- energy-driven superconductivity with a large condensation energy, the electron-doped side is itinerant with a potential- energy-driven superconductivity and a small condensation energy.

The Hubbard Hamiltonian=

0+UDˆ consists of a hop- ping term共“kinetic energy”兲

0= −

i,j,␴

tij共ci

cj+cjci兲 共1兲 and an on-site repulsionUDˆ, where=兺inini is the num- ber of doubly occupied sites, ni=cici, and the operator ci共ci兲creates共annihilates兲an electron at siteiwith spin␴. We use the trial ground state

兩⌿典=e−hHˆ0/te−gDˆ兩⌿0典, 共2兲 wheregandh are共real兲variational parameters and兩⌿0典is a BCS state with d-wave symmetry. The first term in Eq. 共2兲 promotes delocalization and yields a substantial improve- ment with respect to the Gutzwiller wave function共h= 0兲. In fact, it has been demonstrated that ansatz共2兲is very close to the exact ground state both for small 2D systems6and for the solvable 1/rchain.7

In our previous study of the simple Hubbard model 共tij

=t for nearest-neighbor sites and 0 otherwise兲,8 we have found that wave function共2兲exhibitsd-wave pairing below a hole concentration of 0.18. At first sight this result seems to contradict recent Monte Carlo simulations,5where no signa- ture for superconductivity was found, but a closer look shows that there is no discrepancy. In fact, three of the four hole densities considered in Ref.5共0.18, 0.22, 0.28兲are in a region where our variational ground state is also nonsuper- conducting. For the remaining hole density of 0.09 the order parameter共for U/t= 6, the lattice size is 8⫻8, as chosen in Ref.5兲is likely to be too small4to be detected by the simu- lation.

The restriction to nearest-neighbor hopping leads to a 共bare兲Fermi surface which disagrees qualitatively with pho- toemission experiments on layered cuprates.9 Therefore we consider from now on the more realistic case of both nearest- 共t兲 and next-nearest-neighbor hoppings 共t⬘ with canonical parameters U= 8t and t⬘= −0.3t. The bare single-particle spectrum,

k= − 2t共coskx+ cosky兲− 4t⬘coskxcosky,

leads to the Fermi surfaces of Fig. 1. The innermost line corresponds to the Van Hove filling where the Fermi surface passes through the saddle points at 共␲, 0兲 and共0 ,␲兲. These crossings between the Fermi surface and the magnetic zone boundary共the “hot spots”兲 move inward as the densityn is increased and finally merge 共outermost line兲. Hot spots are

kx ky

n = 1.207 n = 1.000 n = 0.726 (0,0)

(0,π)

(π,0) Hot spots

FIG. 1. 共Color online兲Fermi surface for three different electron densities.

Published in "Physical Review B 79(10): 100510(R), 2009"

which should be cited to refer to this work.

http://doc.rero.ch

1

(2)

restricted to 0.726⬍n⬍1.206. Our results, to be discussed below, indicate that superconductivity occurs only in this range.

The BCS state兩⌿0典in Eq.共2兲introduces, in addition tog andh, two other parameters: the amplitude ⌬of the super- conducting gap function and the “chemical potential”␮. To compute the variational energy, the exponent of the operator e−gDˆ is first decoupled by applying a discrete Hubbard- Stratonovich transformation,10 which introduces an Ising spin at each site. All operators are then quadratic in creation and annihilation operators, and therefore the fermionic de- grees of freedom can be integrated out. The remaining sum over Ising spin configurations is performed by a Monte Carlo simulation. In order to avoid the minus sign problem, calcu- lations are carried out in the grand canonical ensemble with an average density fixed by␮, which is therefore not a varia- tional parameter. To reduce the statistical error, the optimiza- tion procedure is based on the method proposed by Ceperley et al.11 and Umrigar et al..12 We have used periodic- antiperiodic boundary conditions.

The minimization of the energyE for fixed average den- sities yields the results of Tables IandIIfor hole and elec- tron dopings, respectively. We notice that the parameter g, which controls the crossover between itinerant共smallg兲and localized 共large g兲 many-particle states, remains large for hole doping but decreases rapidly for electron doping. There- fore, while the hole-doped region 0.75⬍n⬍0.95 is a local- ized doped Mott insulator, the electron-doped part 1.05⬍n

⬍1.2 rapidly undergoes a crossover to an itinerant regime as n increases. 共We estimate a crossover parameter gcr⬇3.兲 This striking difference may root in the bare single-particle density of states at the Fermi energy, which is much larger for hole than for electron doping. We notice that our wave

function is excellent for the itinerant regime and somewhat less reliable for the localized regime.7

Figures 2 and3 show, respectively, the gap parameter⌬ and the superconducting order parameter ⌽=兩具ci

cj

i

典兩 as functions of doping concentration x= 1 −n. The correspond- ing results for t= 0 共Ref. 8兲 are completely electron-hole symmetric and are reproduced only in the right panels. On the hole-doped side superconductivity exists for 0⬍x

⬍0.25共⌬remains finite, but⌽vanishes forx→0兲, i.e., in a larger region than for t⬘= 0. In contrast, on the electron- doped side the superconducting region is reduced to −0.2

x⬍−0.05. Thus we find indeed that superconductivity is restricted to densities where the共bare兲Fermi surface passes through hot spots共see Fig.1兲. Remarkably, for electron dop- ing the gap is suppressed very close to half filling even in the absence of competing antiferromagnetic long-range order.

The marked difference between electron and hole dopings is confirmed by the condensation energy Econd=E共0兲−E共⌬兲, which is 1 order of magnitude smaller for electron doping than for hole doping, as depicted in Fig.4.

TABLE I. Chemical potential, parametersgandh, and energy per site for hole doping共n⬍1兲and an 8⫻8 lattice.

ng h E/t

0.7500 −0.9921共1兲 4.2共1兲 0.113共2兲 −0.858共1兲 0.7800 −0.9612共3兲 4.0共1兲 0.112共2兲 −0.829共1兲 0.8125 −0.9107共3兲 3.9共1兲 0.111共2兲 −0.795共1兲 0.8400 −0.788共1兲 3.7共1兲 0.111共2兲 −0.763共1兲 0.9000 −0.728共1兲 3.8共1兲 0.111共2兲 −0.676共1兲 0.9500 −0.603共1兲 4.0共1兲 0.114共2兲 −0.591共1兲

TABLE II. Chemical potential, parametersgandh, and energy per site for electron doping共n⬎1兲and an 8⫻8 lattice.

ng h E/t

1.0500 0.7666共1兲 3.6共1兲 0.109共2兲 −0.222共1兲 1.0800 0.6440共1兲 3.4共1兲 0.106共2兲 −0.069共1兲 1.1000 0.5488共1兲 3.2共1兲 0.104共2兲 0.040共1兲 1.1300 0.4380共1兲 3.0共1兲 0.100共2兲 0.206共1兲 1.1600 0.3870共1兲 2.9共2兲 0.096共3兲 0.374共1兲 1.2000 −0.2996共1兲 2.6共2兲 0.091共3兲 0.608共1兲

0 0.1 0.2 0 0.3

0.02 0.04 0.06 0.08 0.1 0.12 0.14

t’ = -0.3t, 8x8

0 0.1 0.2 0.3

t’ = -0.3t, 10x10 t’ = 0, 8x8

Hole-doped Electron-doped

Δ/t

x = 1 - n

- - -

FIG. 2.共Color online兲Optimized gap parameter as a function of doping for two different lattice sizes. For comparison, the result for t⬘= 0 is also shown共from Ref.8兲.

0 0.1 0.2 0 0.3

0.01 0.02 0.03 0.04 0.05 0.06

0 0.1 0.2 0.3

t’ = -0.3t, 8x8 t’ = 0, 8x8 Φ= < ci+ci+τ+>

Hole-doped Electron-doped

x = 1 - n

- - -

FIG. 3. 共Color online兲 Superconducting order parameter as a function of doping for an 8⫻8 lattice.

http://doc.rero.ch

2

(3)

Figure5shows the kinetic, potential, and total energies as functions of the gap parameter for the two densities n

= 0.78 and 1.16. For hole doping the energy gain is clearly due to a decrease in kinetic energy, while for electron doping the decrease in potential energy gives the main contribution to the condensation energy. The same behavior has been con- sistently obtained for other densities. The quantity

W= − 2

k

2k

kx

2nk典, 共3兲

also plotted in Fig.5, is—up to a minus sign—proportional to the integrated optical conductivity originating from intra- band transitions.13For the simple Hubbard model 共t= 0W is equal to the kinetic energy, but fort0 the two quantities differ. For hole dopingWhas a pronounced minimum in the region of the optimal gap, corresponding to an increase in oscillator strength, while for electron doping W increases monotonically with⌬, akin to the BCS behavior where the

oscillator strength is reduced at the onset of superconductiv- ity.

We attribute this asymmetry between hole and electron dopings to the different values of the correlation parameterg 共see TablesIandII兲, which puts the hole-doped system into the localized regime, while the electron-doped system is more itinerant. To make the point clear we consider the simple Hubbard model共t⬘= 0兲both in the smallU共itinerant兲 and largeU共localized兲limits. In the smallUlimit supercon- ductivity is produced by the Kohn-Luttinger mechanism,14 where the condensation energy is expected to increase as a function ofU/t, whereas in the largeU limit the condensa- tion energy arises from magnetic exchange and thus is likely to increase with t/U. The change in kinetic energy is then obtained through the Hellman-Feynman theorem,

Ekin共⌬兲−Ekin共0兲= −t

tEcond, 共4兲 which is positive in the small U limit and negative in the largeUlimit.

Additional information can be gained from the magnetic structure factor

Sqជ兲= 1 N

i,j

eiqជ·共Ri−Rj具共nini兲共njnj兲典, 共5兲 displayed in Fig.6.S共qជ兲exhibits a clear maximum at共␲,␲兲, which is largest close to half filling and decreases as doping increases. There is very little difference between electron and hole dopings, presumably because in the large U limit spin correlations depend mostly through the exchange constants J= 4t2/UandJ= 4t2/Uon the microscopic parameters and therefore are essentially independent of the sign oft.

We have seen above that superconductivity is restricted to the region where two points of the Fermi surface can be connected 共at least approximately兲 by the antiferromagnetic wave vector 共␲,␲兲. This together with the strong peak of S共qជ兲for qជ=共␲,␲兲supports the point of view that supercon- ductivity in the two-dimensional Hubbard model is due to a 0

0.1 0.2 0 0.3

0.002 0.004 0.006

0 0.1 0.2 0.3 0

0.01 0.02 0.03

t’ = -0.3t t’ = 0

Electron-doped Hole-doped

- - -

x = 1 - n Econd/t

FIG. 4. 共Color online兲 Condensation energy per site for an 8⫻8 lattice.

0 0.01 0.02 0.03 Δ/t -1

-0.5 0 0.5 1 1.5

0 0.01 0.02 0.03 Δ/t

-2 -1 0 1 2 3

Ekin/t Epot/t E/tW/t

Hole-doped Electron-doped

n = 0.78

n = 1.16 10-3

x x10-3

FIG. 5. 共Color online兲 Changes in kinetic, potential, and total energies as well as in the quantityWas functions of the gap param- eter for an 8⫻8 lattice. 共−W is proportional to the oscillator strength for intraband transitions.兲 The relative uncertainties are smaller than the symbol sizes.

0 1 2 3 4 5

S(q)

0.050.10 0.160.20

0 1 2 3 4 5

0.050.10 0.160.22 0.25

(0,0) (π,π)

(π,0) (0,0)

Electron-doped

(0,0) (π,0) (π,π)

Hole-doped

FIG. 6. 共Color online兲Magnetic structure factor as a function of the momentum, for various densities above and below half filling, for an 8⫻8 lattice.

http://doc.rero.ch

3

(4)

magnetic mechanism. This conclusion is now widely ac- cepted, but the question whether the mechanism is of the resonating valence bond 共RVB兲type1 or arises from the ex- change of spin fluctuations15 is presently under debate.16,17 We cannot solve this problem on the basis of our variational calculations, but the distinction between localized and itiner- ant regimes can give additional useful insight. According to our results the former is appropriate for the hole-doped re- gion, where superconductivity is associated with a decrease in kinetic energy; the latter regime is found for electron dop- ing, where superconductivity arises from a more conven- tional gain in potential energy.

The results described above compare surprisingly well with experiments on cuprates, better than our earlier calculations8where only nearest-neighbor hopping has been taken into account. This concerns the phase diagram, in par- ticular, for electron-doped materials for which a recent sys- tematic study18gives a dome shape共forTc兲strikingly similar to our Fig. 2. Photoemission data give values of the super- conducting gap ⌬ in the range 10–20 meV for hole-doped compounds 关La2−xSrxCuO4+y 共LSCO兲 共Ref. 19兲 or Bi2201 共Ref. 9兲兴 and ⬃5 meV for an electron-doped compound 关NdCeCO共Ref. 20兲兴. Choosingt= 300 meV 共neutron data兲,

our maximum gap parameters are 15 and 7 meV for hole and electron dopings, respectively. An increase in oscillator strength, predicted by our calculations for hole doping, has been reported on the basis of optical spectroscopy for an underdoped sample,21 but the situation is less clear on the overdoped side. We are not aware of corresponding measure- ments on electron-doped materials.

In conclusion, our variational search ford-wave supercon- ductivity in the two-dimensional Hubbard model gives a dif- ferentiated picture, namely, a large共moderate兲correlation pa- rameter for hole 共electron兲 doping, a gain in kinetic 共potential兲energy due to pairing, and a large共small兲conden- sation energy. Our wave function is expected to be better suited for describing the electron-doped region, which is less localized than the hole-doped region. A more elaborate study is needed to improve our understanding of the pseudogap phase observed in the cuprates for weak hole doping.

We are grateful for financial support from the Swiss Na- tional Science Foundation through the National Center of Competence in Research “Materials with Novel Electronic Properties-MaNEP.” The computations have been partially performed on the Pléiades cluster of the EPFL.

1P. W. Anderson, Science 235, 1196共1987兲.

2T. A. Maier, M. Jarrell, and D. J. Scalapino, Phys. Rev. B 74, 094513共2006兲.

3K. Haule and G. Kotliar, Phys. Rev. B 76, 104509共2007兲.

4H. Yokoyama, Y. Tanaka, M. Ogata, and H. Tsuchira, J. Phys.

Soc. Jpn. 73, 1119共2004兲.

5T. Aimi and M. Imada, J. Phys. Soc. Jpn. 76, 113708共2007兲.

6H. Otsuka, J. Phys. Soc. Jpn. 61, 1645共1992兲.

7M. Dzierzawa, D. Baeriswyl, and M. Di Stasio, Phys. Rev. B 51, 1993共1995兲.

8D. Eichenberger and D. Baeriswyl, Phys. Rev. B 76, 180504共R兲 共2007兲.

9A. Damascelli, Z. Hussain, and Z.-X. Shen, Rev. Mod. Phys. 75, 473共2003兲.

10J. E. Hirsch, Phys. Rev. B 28, 4059共1983兲.

11D. Ceperley, G. V. Chester, and K. H. Kalos, Phys. Rev. B 16, 3081共1977兲.

12C. J. Umrigar, K. G. Wilson, and J. W. Wilkins, Phys. Rev. Lett.

60, 1719共1988兲.

13D. Baeriswyl and L. Degiorgi, Strong Interactions in Low Di- mensions: Physics and Chemistry of Materials with Low- Dimensional Structures共Kluwer, Dordrecht, 2004兲, Vol. 25.

14W. Kohn and J. M. Luttinger, Phys. Rev. Lett. 15, 524共1965兲.

15D. Scalapino, J. Low Temp. Phys. 117, 179共1999兲.

16P. W. Anderson, Science 316, 1705共2007兲.

17T. A. Maier, D. Poilblanc, and D. J. Scalapino, Phys. Rev. Lett.

100, 237001共2008兲.

18Y. Krockenberger, J. Kurian, A. Winkler, A. Tsukada, M. Naito, and L. Alff, Phys. Rev. B 77, 060505共R兲 共2008兲.

19K. Yamada, S. Wakimoto, G. Shirane, C. H. Lee, M. A. Kastner, S. Hosoya, M. Greven, Y. Endoh, and R. J. Birgeneau, Phys.

Rev. Lett. 75, 1626共1995兲.

20T. Sato, T. Kamiyama, T. Takahashi, K. Kurahashi, and K. Ya- mada, Science 291, 1517共2001兲.

21F. Carboneet al., Phys. Rev. B 74, 064510共2006兲.

http://doc.rero.ch

4

Références

Documents relatifs

scales and the distance to the threshold. The derivation of amplitude equations at orders higher than 83/2 starting with the Boussinesq equations is very painful and it

Fig.l.-Dotted line : Fermi surface for a square half filled tight binding model with nearest neighbour overlap

Table 2 Functional groups (1–4) organized by genera, trophic preference and habitat Genera Family Troph ic H abitat G ID Gene ra Fam ily Troph ic H abitat G ID Mant a MYLIO BATIDAE

The two described algorithms, based on optimal job insertion (using job inser- tion algorithm OJI-NWJS) and local search with neighborhoods N 1 and N 1 ∪ N 2 respectively, will

For the square, triangular, and honeycomb lattices, we provide detailed superconducting phase diagrams as well as coupling strengths which quantify the corresponding critical

These loop models exhibit different critical behaviour than the standard FPL model on the triangular lattice, which is in the universality class of dense loops [18, 20].

8, 9, 10 and 11 show the DC conductivity, the structure factor and the double occupancy curves for different values of the staggered potential ∆ at kT = 0.01.. For ∆ &amp; 1 the

Our analysis builds on a 2D lattice model with two interaction scales: A positive on-site Coulomb interaction U, which may be tuned to target the regime of strong